You are on page 1of 35

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/263702407

Polyoxometalate-Supported Lanthanoid
Single-Molecule Magnets
ARTICLE in AUSTRALIAN JOURNAL OF CHEMISTRY JULY 2014
Impact Factor: 1.64 DOI: 10.1071/CH14166

VIEWS

41

2 AUTHORS:
Michele Vonci

Colette Boskovic

University of Melbourne

University of Melbourne

3 PUBLICATIONS 70 CITATIONS

69 PUBLICATIONS 1,459 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Michele Vonci


Retrieved on: 25 July 2015

Polyoxometalate-Supported Lanthanoid Single-Molecule Magnets

Michele Vonci,A Colette BoskovicA,B


A

School

of

Chemistry,

The

University

of

Melbourne,

Vic

3010,

Australia.

Corresponding author. Email: c.boskovic@unimelb.edu.au

Polyoxometalates are robust and versatile multidentate oxygen-donor ligands,


eminently suitable for coordination to trivalent lanthanoid ions. To date ten very
different structural families of such complexes have been found to exhibit slow
magnetic relaxation due to single-molecule magnet (SMM) behaviour associated with
the lanthanoid ions. These families encompass complexes with between one and four of
the later lanthanoid ions: Tb, Dy, Ho, Er and Yb. The lanthanoid coordination numbers
vary between six and eleven and a range of coordination geometries are evident. The
highest energy barrier to magnetisation reversal measured to date for a lanthanoidpolyoxometalate SMM is Ueff/kB = 73 K for the heterodinuclear Dy-Eu compound
(Bu4N)8H4[DyEu(OH)2(-SiW10O36)2].

Introduction

The search for molecular materials as an alternative or complement to bulk materials is


driven by practical demands imposed by technological progress, as well as by advances in
fundamental chemistry and physics that facilitate the synthesis and investigation of such
species. Examples of the molecular materials approach are liquid crystals, organic

conductors and semi-conductors and organic luminescent materials. Molecular magnets


represent the same technological and conceptual development in the field of magnetism.
In this regard single molecule-magnets (SMMs) are molecular species that can retain
magnetisation at low temperature, due to an anisotropy-induced energy barrier to
magnetisation reversal.[1-3] Single-molecule magnets are bistable at low temperature,
which implies potential for applications in high density data storage. Magnetisation
quantum tunnelling is also a feature of SMMs, which gives rise to quantum states that
consist of coherent superpositions of the "spin-up" and "spin-down" states. The quantum
tunnelling in SMMs can be switched on and off by the application of small magnetic
fields, which may ultimately provide the superposition of logic states necessary for SMMs
to act as qubits in quantum computers.[4]
Common experimental techniques to evaluate the slow magnetic relaxation in
SMMs are alternating field (ac) susceptibility, magnetisation hysteresis and
magnetisation relaxation measurements. The presence of a frequency-dependent out-ofphase component of the ac magnetic susceptibility (

m'')

indicates slow magnetic

relaxation and is a key characteristic of a SMM. The presence of peak maxima in the outof-phase susceptibility allows the extraction of the magnetisation relaxation time of the
system at different temperatures and frequencies. The assumption of an Arrhenius-like
dependence of the relaxation time on the temperature allows determination of the
energy barrier for the thermally activated relaxation process (Ueff). However, other
relaxation paths can coexist with the thermally activated one, the most relevant of which
is quantum tunnelling, which particularly affects lanthanoid-based SMMs (Ln-SMMs).[5]
The presence of quantum tunnelling can impede the observation of hysteresis or the
detection of peak maxima in the out-of-phase ac susceptibility. In the latter case a

common practice is the application of a weak dc magnetic field to suppress quantum


tunnelling during ac susceptibility measurements, thus allowing evaluation of the energy
barrier for the thermally activated process. Some controversy surrounds this practice and
the interpretation of the resulting data. In a recent review, Layfield and co-workers
suggested that single-molecule magnetism should only be claimed in the case of clear
frequency-dependent peak maxima in the out-of phase ac susceptibility or the presence
of magnetisation hysteresis.[6] These authors also suggest that the magnitude of any
external dc field used to estimate the energy barrier of the thermally activated process
should be explicitly stated, in order to facilitate comparison between species.
Following the discovery of single-molecule magnetism in the early 1990s, for
some years researchers focused on exchange-coupled polynuclear complexes of 3d
transition metals, especially manganese(III). The zero-field splitting of the "giant spin"
ground state in these complexes gives rise to an energy barrier to magnetisation reversal
and therefore slow magnetic relaxation at low temperatures. To date the highest energy
barrier

reported

for

polynuclear

SMMs

are

Ueff/kB

86.4

for

[MnIII6O2(sao)6(O2CPh)2(EtOH)4] (saoH2 = salicylaldoxime) and Ueff/kB = 96 K for


[Co(hfpip)2(D2py2(TBA))]2 (hfpip = 1,1,1,5,5,5-hexafluoro-4-(4-tert-butylphenylimino)-2pentanonate;

D2py2(TBA)

4,4'-(5-(3,3-dimethylbut-1-ynyl)-1,3-

phenylene)bis(diazomethylene)dipyridine).[7,8] After considerable experimental effort in


this direction, the realisation that increasing the ground state spin of polynuclear
complexes is not an efficient strategy for achieving a large energy barrier to
magnetisation relaxation,[9] led to the examination of mononuclear complexes of highly
anisotropic transition metals as SMMs. The highest energy barrier reported for a
mononuclear SMM based on a 3d metal is Ueff/kB = 325 K for the mononuclear iron(I)

compound, K(crypt-222)][Fe(C(SiMe3)3)2] (crypt-222 = 2.2.2-cryptand).[10] Considerable


impetus to the investigation of lanthanoid rather than d-block metals for developing new
and better SMMs arose following the 2003 report of SMM behaviour for "double-decker"
lanthanoid-phthalocyanine (Pc) compounds (Bu4N)[LnPc2] (Ln = Tb, Dy). These
compounds display slow magnetic relaxation at higher temperatures (by ac susceptibility)
than has ever been observed for transition metal based SMMs and with much larger
energy barriers for magnetisation reversal.[11,12] The present record for a lanthanoidbased SMM is Ueff/kB = 938 K for the heteroleptic bis(phthalocyanine) complex [Tb{(O(C6H4)-p-tBu)8Pc}(Pc)].[13]
It has become clear that molecules containing one or more lanthanoid ions can
exhibit slow magnetic relaxation of molecular origin and behave as SMMs.[14-17] They are
sometimes referred to as single-ion magnets (SIMs) because the energy barrier to
magnetisation reversal arises from single ion properties of the lanthanoid(III) ions, rather
than from zero-field splitting of a "giant spin" ground state in an exchange-coupled
polynuclear complex. The magnetic bistabilty of Ln-SMMs results from the energy level
structure that arises when the spin orbit coupled ground state J term is further split
under the effect of the crystal field into 2J + 1 sublevels. The precise nature of the
splitting depends on the symmetry and on the Kramers or non-Kramers structure of the
ion.[18] An easy axis with preferential orientations of the magnetisation is then possible
when substates with high values of MJ are stabilised.
Rinehart and Long have established a simple qualitative approach for the
rationalisation of the magnetic behaviour of Ln-SMMs in terms of oblate or prolate
overall quadrupole charge moment of the trivalent lanthanoid ions in relation to the
surrounding ligand arrangement when an easy axis of the magnetisation is present.[19]

Limiting the analysis only to the late lanthanoids with high magnetic moment J, the
oblate charge distribution of ions such as Tb(III), Dy(III), and Ho(III) is stabilised by the
negative charges of the ligands in axial positions, whereas the prolate charge distribution
of Er(III), Tm(III), and Yb(III) is stabilised by an equatorial arrangement of the ligands. The
model describes the non-sequential ordering often found in the relative energies of MJ
substates, taking into account the quadrupolar charge disposition of the different MJ
terms arising from the crystal field splitting of the ground state term, instead of the
interaction of the overall charge with the ligands. This approach can be useful both for
prediction and for a-posteriori magnetostructural correlation.[20-23]
Since the discovery of SMM behaviour of (Bu4N)[LnPc2] complexes, more than 160
Ln-SMMs have been reported,[6] spanning from mononuclear to polynuclear molecules
with various different classes of ligands. Efforts have focused on the trivalent ions of the
later lanthanoids Tb, Dy, Ho and Er, the ground states of which have large total magnetic
moment and the requisite magnetic anisotropy. By far the most examples of Ln-SMMs
are based on Dy and Tb, due to their large magnetic anisotropy and total magnetic
moment.[6] A number of polynuclear lanthanoid complexes exhibit SMM behaviour,
however in general exchange coupling has not been found to play a significant role, with
the SMM behaviour arising solely from single ion properties.

Polyoxometalates
Polyoxometalates (POMs) are molecular fragments of early transition metal oxides
formed from acid condensation of simple oxoanions.[24-26] They are comprised of metalcentered MOn polyhedra (M is most commonly Mo, W or V) linked by shared corners,
edges or faces. These are referred to as heteropolyanions when additional heteroatoms

(eg P, As, Si and Ge) are incorporated within the coordination cluster, otherwise species
without heteroatoms are referred to as isopolyanions. The reversible and multi-electron
redox chemistry that POMs can display confers an ability to access multiple electronic
states through chemical, electrochemical or photophysical processes.[27,28]
In their typical plenary closed shell form, POMs have little reactivity. This can be
enhanced by base hydrolysis to obtain more reactive defect, or lacunary, species (Fig. 1).
Structural isomerism is a common phenomenon of POMs, which is also evident for the
lacunary forms. The oxygen-rich surface of lacunary POMs renders them excellent
multidentate inorganic ligands for oxophilic d- and f-block metals (Fig. 1), giving rise to a
wide variety of coordination complexes. This includes numerous lanthanoid complexes
with POM ligands (Ln-POMs), which are of interest for their magnetic, luminescent,
catalytic and electronic properties. [29-33]

Fig. 1 - Ball and stick and polyhedral representations of some common POM structural
families and their lacunary derivatives employed in Ln-POM SMMs: a) Lindqvist, b)
Keggin, c) Dawson, d) Preyssler and e) Anderson. Colour code: Mo/W, yellow or grey; O,
red; heteroatom, blue.

Lanthanoid-Polyoxometalate Single-Molecule Magnets

Inorganic POMs offer several advantages over organic species as supporting ligands for
Ln-SMMs. Due to the steric bulk and intrinsic diamagnetism of the oxidised forms, they
can ensure good magnetic insulation of the lanthanoid ions in neighbouring molecules,
minimising intermolecular interactions that can give rise to unwanted quantum
tunnelling and loss of magnetisation. In addition, the magnetic behaviour can be ascribed
to single-molecule phenomena and not to long range intermolecular interactions. As
ligands POMs can offer a variety of local coordination environments to provide different
crystal fields to the lanthanoid ion, some of which are not accessible with organic ligands.
To date ten structural families of lanthanoid complexes with POM ligands have been
reported to exhibit slow magnetic relaxation consistent with SMM behaviour (Table 1)
and a variety of lanthanoid coordination geometries are evident in these complexes (Fig.
2). These compounds are grouped in the discussion that follows according to the number
of lanthanoid centres in each complex.

Table 1. Relevant parameters for reported lanthanoid-polyoxometalate single-molecule


magnets
Formula

Abbreviation

Ueff/kb (K)
(Hdc (Oe))

Na9[Ln(W5O18)2]

Hysteresis

Ref

(K)

1-Ho

36

1-Er

55.2

36

2-Dy

36

2-Ho

36

2-Er

36

2-Yb

36

3-Dy

24

38

3-Ho

0.8

38

4-Dy

11.8 (1000)

42

5-Dy2

0.8-1.6

43

6-DyEu

73

44

6-Dy2

65.7

43

6-DyYb

53 (1800)

44

6-DyLu

48

44

[{CuTbL(H2O)3}2{IMo6O24}]Cl

7-Tb

17.1

45

[{CuTbL(H2O)2}2{AlMo6O18(OH)6}2]

8-Tb

20.8

45

K8nH3n[Dy3nKn(H2O)3(CO3)(A--AsW9O34)(A--

9-Dy

54.3 (5000)

46

10-Dy

3.9

47

K13[Ln(2-SiW11O39)2]

K12[LnP5W30O110]

(Me4N)8[Ln(H2O)8]2
[Ln2(H2O)4(2-As2W17O61)2]
(Bu4N)6H4
[{Ln(H2O)2(CH3COCH3)}2(-SiW10O36)2]
(Bu4N)8H4
[Ln1Ln2(OH)2(-SiW10O36)2]

AsW9O34)] (n = 0 or 1)
(HDABCO)8H5Li8[Ln4As5W40O144(H2O)10(gly)2]
a

L = N,N-bis(3-methoxysalicylidene)-ethylenediamine, DABCO = 1,4-diazabicyclooctane, gly = zwitterionic


glycine.
b
Ueff values are obtained from measurements in zero applied dc field, except as indicated in parentheses

10

Fig. 2. Approximate coordination geometries for reported Ln-POM SMMs:[34] a) sixcoordinate: trigonal prismatic (6); b) seven-coordinate: capped trigonal prismatic (5, 9),
capped octahedral (4); c) eight-coordinate: square antiprismatic (1, 2, 10); trigonal
dodecahedral (10); biaugmented trigonal prismatic (8); d) nine-coordinate: muffin (7); e)
eleven-coordinate: capped pentagonal antiprismatic (3). Colour code: O, red; Ln, violet.

Mononuclear {Ln1} Compounds


To date three structural families of mononuclear Ln-POM SMMs have been reported:
Na9[Ln(W5O18)2] (Ln = Tb, Dy, Ho, and Er) (1),[35-37] K13[Ln(2-SiW11O39)2] (Ln = Tb, Dy, Ho,
Er, Tm, and Yb) (2) [36] and K12[LnP5W30O110] (Ln = Tb, Dy, Ho, Er, Tm, and Yb) (3) [38] (Fig.
3). Known for many years, members of family 1 are obtained from the reaction of
lanthanoid ions with the monolacunary form of the Lindqvist POM generated in situ by
acid condensation of sodium tungstate.[39] Magnetic studies were performed on the
sodium salts. A similar reaction between lanthanoid ions and the potassium salt of the

11

monolacunary Keggin [2-SiW11O39]8- ion affords family 2.[39] Family 3 complexes are
obtained as potassium salts following hydrothermal reaction between a mixed salt of the
Preyssler POM K12.5Na1.5[NaP5W30] and lanthanoid chloride.[40,41]

Fig. 3. Structural representation of the polyanions of mononuclear Ln-POM SMMs (a)


[Ln(W5O18)2]9- (1), (b) [Ln(2-SiW11O39)2]13- (2), (c) [LnP5W30O110]12 (3). Colour code as per
Fig. 1.

Two monolacunary {W5O18} Lindqvist or {2-SiW11O39} Keggin POMs act as


tetradentate oxygen donor ligands, effectively "sandwiching" the 8-coordinate
lanthanoid centres in 1 and 2, respectively. The local coordination geometries of the
lanthanoid centres in 1 and 2 are distorted square antiprismatic with approximate D 4d
point symmetry (Fig. 2).[34] Axial compression of the square antiprism is evident in both
complexes, with 2 also exhibiting a rhombic distortion. For 3 the 11-coordinate
lanthanoid centre is encapsulated within the [P5W30O110]15 Preyssler POM, in a site

12

otherwise occupied by alkali cations. The coordinate geometry is approximately capped


pentagonal antiprismatic,[34] with approximate C5 point symmetry (Fig. 2). Of the eleven
oxygen centres that bind to the lanthanoid, five bridge to tungsten centres, five bridge to
phosphorus centers and one is a terminal aqua ligand.
Variable temperature magnetic susceptibility measurements of 1, 2, and 3 afford
room temperature

mT

values in good agreement with those expected for the free ions

(Fig. 4).[36,38] Decreases in the

mT

values with temperature are ascribed to crystal field

effects. For series 1, ac susceptibility measurements in the range 2-10 K with frequencies
in the range 1-10,000 Hz reveal SMM behaviour for 1-Ho and 1-Er. For 1-Ho a shoulder is
present in

m''

with no clear maxima, which is strongly frequency-dependent and tends

to disappear when the frequency is reduced. Arrhenius analysis was not carried out and
the energy barrier for the reversal of magnetisation was not estimated. Compound 1-Er
exhibits a strong frequency dependence and clear maxima for

m''

(Fig. 4). Arrhenius

analysis yields a barrier height (Ueff/kB) of 55.2 K (pre-exponential factor, 0 of 1.6 10-8
s). The temperature dependence of

m''

shows two relaxation regimes. Relaxation is

thermally activated above 3 K, while below 3K it is essentially temperature independent,


due to a tunnelling process between the ground state. For series 2 ac susceptibility
measurements show frequency-dependent behaviour for 2-Dy, 2-Ho, 2-Er and 2-Yb but
no maxima were evident in

and it was not possible to determine Ueff.

13

Fig. 4. Plot of
(

m'')

mT

vs T for members of family 1 (top). In phase (

m')

and out-of-phase

ac susceptibility data for 1-Er for frequencies in the range 1-10 KHz (bottom).

(These images are adapted with permission from ref. 36. Copyright 2009 The American
Chemical Society)

For series 3 ac susceptibility data indicate SMM behaviour for 3-Dy and 3-Ho.[38]
For 3-Dy both

m'

and

m''

are frequency dependent but do not show clear maxima above

80 mK, with little dependence of relaxation time upon temperature. Nevertheless, by


estimating the relaxation time as =

m/( m)

it was possible extract the value of the

energy barrier to the reversal of the magnetisation Ueff/kb = 24 K (0 not reported) for the
thermally activated process above 2 K. Below 5 K the relaxation starts to become

14

temperature-independent with the quantum tunnelling driven relaxation process


switching on at 2 K. Frequency-dependent maxima in

m''

plot were observed at very low

temperature for 3-Ho. Arrhenius analysis afforded the very small energy barrier to
magnetisation reversal of Ueff/kb = 0.8 K (0 = 6 103 s). For both 3-Dy and 3-Ho
magnetisation hysteresis loops observed below 2 K provide further evidence of SMM
behaviour (Fig. 5). The presence of hysteresis despite the fast quantum tunnelling
mechanism is ascribed in the quantum tunnelling regime for both compounds to the
increase in relaxation time in presence of an applied magnetic field.

Fig. 5. Magnetisation hysteresis loop for 3-Dy (top) and 3-Ho (bottom). (This image is
adapted from ref. 38 with permission. Copyright 2012 The American Chemical Society.)

15

Following the procedure described by Ishikawa for the double-decker


phthalocyanine complexes,[12] the

mT

versus T data for each of the different analogues in

the isostructural series 1, 2, and 3 were simultaneously fit to a crystal field Hamiltonian.
This afforded the energies and the corresponding wavefunctions arising from the splitting
of the J multiplet of the lanthanoid ion ground states. The Hamiltonian was simplified to
take into account the pseudo-symmetry of the coordination sphere of the lanthanoid ion,
allowing reduction of the number of free parameters. Qualitatively similar energy level
diagrams were obtained for the corresponding analogues of families 1 (Fig. 6) and 2,
which is not surprising considering the assumption of D4d pseudo-symmetry in both
cases. The C5 point symmetry assumed for 3 afforded a different pattern, with this
symmetry requiring considerable mixing of the levels. For 1-Er and 1-Ho, EPR and specific
heat measurements confirm the relatively large ground state MJ values determined from
the magnetic susceptibility,[36,37] which are consistent with SMM behaviour observed for
these compounds. Despite the similar energy pattern determined for 2, the lack of
maxima in the

m''

plots are likely due to the less than axial symmetry giving rise to

rhombic contributions to the magnetic anisotropy and a significant admixture of states


which enhances magnetic relaxation via quantum tunnelling. For 3-Er calculated
admixture of the ground and first excited state explains the absence of slow magnetic
relaxation. In contrast, the ground states for both 3-Dy and 3-Ho are characterised by
relatively pure Kramers doublets with the highest possible |MJ| values, well separated
from the first excited states. This situation is favourable for SMM behaviour, consistent
with experimentally observed slow magnetic relaxation. The Ueff value experimentally
determined for 3-Ho is significantly smaller than the energy gap between the ground
sublevel to the first excited level. This is attributed to a relaxation process driven by a

16

hyperfine interaction in which magnetic relaxation takes place by thermally activated


transitions to excited nuclear spin states. Finally, terbium analogues of all three
compounds 1-Tb, 2-Tb and 3-Tb exhibit non-degenerate ground states, which implies no
possibility of SMM behaviour, consistent with experiment.

Fig. 6. Energy levels and MJ values calculated for series 1. (This image is adapted with
permission from ref. 36. Copyright 2009 The American Chemical Society)

Dinuclear {Ln2} Compounds


Five structural families of dinuclear Ln-POM complexes (Fig. 7) have shown evidence of
slow

magnetisation

relaxation:

(Me4N)8[Ln(H2O)8]2[Ln2(H2O)4(2-As2W17O61)2]

(4),

(Bu4N)6H4[{Ln(H2O)2(CH3COCH3)}2(-SiW10O36)2] (5), (Bu4N)8H4[Ln1Ln2(OH)2(-SiW10O36)2]


(6), [{CuTbL(H2O)3}2{IMo6O24}]Cl (7) and [{CuTbL(H2O)2}2{AlMo6O18(OH)6}2] (8) (L = N,Nbis(3-methoxysalicylidene)-ethylenediamine for 7 and 8).[42-45]
Dysprosium (4-Dy) and erbium (4-Er) analogues of family 4 are obtained as
tetramethylammonium salts following reaction of the trilacunary Dawson POM [-

17

As2W15O56]12- with lanthanoid ions, which must involve partial dissociation and
rearrangement of the POM to afford the monlacunary dawson unit evident in the
product.[42] Mixed tetrabutylammonium salts of compound 5-Ln2 (5-Gd2, 5-Dy2) are
synthesised from the reaction in acetone of dilacunary Keggin POM [-SiW10O36]8- with
Dy(acac)3.[43] Homodinuclear compounds 6-Ln2 can be obtained by recrystallisation of 5Ln2 from dichloromethane in presence of tetrabutylammonium hydroxide and 6-Ln2 can
be converted back to 5-Ln2 by recrystallisation from acidified acetone/acetonitrile/water
mixture in presence of diethyl ether.[43] The synthesis of heterodinuclear 6-DyLn
complexes involves isolation of a mononuclear species by mixing Dy(acac)3 and TBA4H4[SiW10O36] in acetone/water mixture followed by the addition of diethyl ether. [44] The final
selective formation of the heterodinuclear compound is obtained by adding Ln(acac)3 to
a dichloromethane solution of the monovacant 6-Dy species. Compounds 7 and 8 are
synthesised by layering a methanol solution of the precursor complex [CuTbL(H2O)3Cl2]+
with an aqueous solution of the A-Anderson POM [IMo6O24]5- (for 7) or the B-Anderson
POM [AlMo6O24H6]3- (for 8).[45]

18

Fig.

7.

Structural

representations

of

the

polyanions

in

a)

(Me4N)8[Ln(H2O)8]2[Ln2(H2O)4(2-As2W17O61)2] (4), b) (Bu4N)6H4[{Ln(H2O)2(CH3COCH3)}2(SiW10O36)2]

(5),

c)

(Bu4N)8H4[Ln1Ln2(2-OH)2(-SiW10O36)2]

(6),

d)

[{CuTbL(H2O)3}2{IMo6O24}]Cl (7) and e) [{CuTbL(H2O)2}2{AlMo6O18(OH)6}2] (8). Colour code:


as per Fig. 1; Cu, cyan.

The polyanions in compounds 4, 5 and 6 are all comprised of two lanthanoid


centres sandwiched between two monolacunary dawson POM ligands (4) or dilacunary
Keggin POM ligands (5 and 6). The lanthanoid centres (Fig. 2) are: 7-coordinate distorted
monocapped octahedral (4), 7-coordinate distorted monocapped trigonal prismatic (5)
and 6-coordinate distorted trigonal prismatic (6).[34] The intramolecular LnLn distances

19

are 6.26, 4.49 and 3.60-3.65 in 4, 5 and 6, respectively. Two -hydroxo ligands directly
bridge the lanthanoid centres in 6. Notably, compound 4 incorporates lanthanoid ions
additional to those with POM anions as [Ln(H2O)8]3+ countercations linked to the
polyanions by hydrogen bonding.
Compounds 7 and 8 are structurally quite different from 4-6. They both
incorporate two terbium centres bridged by one (7) or two (8) Anderson POM ligands,
with each terbium centre also coordinated to a copper atom via a hexadentate (N 2O4)
organic ligand. The terbium centres are 9-coordinate "muffin" (7) and 8-coordinate
biaugmented trigonal prismatic (8). [34]
The temperature dependence of the ac susceptibility (2-15 K, 100-1500 Hz) was
measured for 4-Er and 4-Dy. In absence of an external dc magnetic field frequencydependent

m''

is evident only for 4-Dy, although peak maxima were not evident above 2

K.[42] However in presence of an external static magnetic field of 1000 Oe and with an
oscillating magnetic field of 3 Oe (1001500 Hz), frequency-dependent

m''

a is observed

for both 4-Er (no peak maxima) and 4-Dy (peak maxima evident above 2 K). The
Arrhenius analysis for 4-Dy results in a Ueff of 11.8 K (

= 1.4 10-6 s).

The magnetic susceptibility data of the spin-only analogues 5-Gd2 and 6-Gd2 were
measured to investigate the intramolecular interaction between the two lanthanoid
ions.[43] The data were fit to an isotropic Heisenberg-Dirac-Van Vleck model (H = 2JS1S2)
resulting in g = 2.00 and J = 0.002 cm-1, for 5-Gd2 and g = 2.00 and J = -0.22 cm-1 for 6-Gd2,
showing that the bis(2-OH) bridge mediates a larger, although still modest,
intramolecular interaction between lanthanoid ions. Alternating current susceptibility
measurements reveal a dependency of

m''

on frequency for 5-Dy2, although no peaks

were observed above 2 K. [43] The energy barrier to magnetisation reversal was estimated

20

to be 0.8-1.6 K (

in the range 7.6 10-6-5.2 10-5 s). In the absence of an applied dc

field, ac susceptibility measurements for 6-Dy2 reveal frequency dependent


with clear maxima evident in

m''

m'

and

m'',

(Fig. 8). The Arrhenius analysis indicates a thermally

activated process above 4 K, with Ueff of 65.7 K (

= 3.11 10-7 s). Fast relaxation due to

quantum tunnelling is evident below 4 K. The role of single ion anisotropy in SMM
behaviour of 6-Dy2 was investigated by synthesising a sample magnetically diluted with
Y(III) 6-DyY (a statistical mixture of 1 % DyDy, 18 % DyY and 81 % YY).[43] The Arrhenius
analysis on the diluted product yielded similar results to those of 6-Dy2 with Ueff/kb = 62.6
K (0 = 4.2 10-7 s), showing that the intramolecular DyDy magnetic interaction has
little effect and that the main source of SMM behaviour is the single ion anisotropy of the
dysprosium(III) ion. The magnetic behaviour of heterodinuclear analogues of 6 are also of
interest, with 6-DyEu and 6-DyLu exhibiting frequency dependent

m''

and clear maxima

above 2 K.[44] Peak maxima were only observed for 6-DyYb upon application of an
external dc field (1800 Oe) to suppress quantum tunnelling. The relevant parameters for
the three systems were extracted (Fig. 9) resulting in Ueff/kb = 73 K (0 = 1.8 10-7 s) for 6DyEu, Ueff/kb = 53 K (0 = 8.2 10-7 s) for 6-DyYb, and Ueff/kb = 48 K (0 = 1.9 10-6 s) for 6DyLu. The value of the energy barrier of the series increases in the order of increasing
ionic radius: 6-DyLu (48 K) < 6-DyYb (53 K) < 6-Dy2 (66 K) < 6-DyEu (73 K) (Fig. 9), which is
ascribed primarily to the effect of changing the local coordination of the dysprosium(III)
ion.

21

Fig. 8. In-phase and out-of-phase ac susceptibility of 6-Dy2 in the range of frequencies


0.71500 Hz. (This image is adapted from ref. 43 with permission of The Royal Society

of Chemistry.)

Fig. 9. Arrhenius plots of 6-DyEu (), 6-DyYb (), and DyLu (). The solid lines represent
the best fits to the Arrhenius equation under the thermally activated regime. (This image
is adapted from ref. [44] with permission. Copyright 2013 Wiley-Vch Verlag GmbH & Co.
KGaA, Weinheim)

22

For both 7-Tb and 8-Tb the value of

mT

at room temperature is in good

agreement with the expected value for two independent copper(II) ions and two
terbium(III) ions.[45] The behaviour of

mT

at lower temperatures shows complex

concomitant phenomena of depopulation of states and magnetic anisotropy for


terbium(III) ion and a ferromagnetic interaction with the copper centre. The ac
susceptibility measurements for both 7-Tb and 8-Tb exhibit clear frequency-dependent
maxima

of

m'',

thus

indicating

SMM

behaviour.

The

precursor

complex

[CuTbL(H2O)3Cl2]ClMeOH does not exhibit slow magnetic relaxation, and this change in
magnetic behaviour is attributed to the different coordination environments provided to
the terbium(III) ion by the POM ligands. The relaxation behaviour exhibits two distinct
temperature regimes: thermally activated above 4 K and another temperatureindependent below 4K due to fast quantum tunnelling of magnetisation. For data in the
thermally activated regime Arrhenius analysis yielded an energy barrier of 17.1 K (0 = 4.0
106) and 20.8 K (0 = 1.1 106) for 7-Tb and for 8-Tb, respectively.

Trinuclear {Ln3} Compounds


The disordered Dy-POM compound K8nH3n[Dy3nKn(H2O)3(CO3)(A--AsW9O34)(A-AsW9O34)] (n = 0 or 1) (9-Dy) exhibits SMM behaviour.[46] The major component
polyanion is a trinuclear dysprosium species of formula [Dy3(H2O)3(CO3)(A--AsW9O34)(A-AsW9O34)]11 (Fig. 10). Compound 9-Dy is synthesised from the reaction of the
trilacunary Keggin POM {A--AsW9O34) with dysprosium ions in a slightly basic solution in
the presence of citric acid, which acts as a source of carbonate ion. The major polyanion
component of 9-Dy is comprised of a triangular arrangement of dysprosium centres

23

bridged by a

3-carbonate

ligand and sandwiched between two different trilacunary

Keggin POMs. The dysprosium centres are 7-coordinate with distorted monocapped
trigonal prismatic coordination geometry.[34] The DyDy distances are in the range 4.855
to 4.863 .

Fig. 10. Structural representation of the polyanion [Dy3(H2O)3(CO3)(A--AsW9O34)(A-AsW9O34)]11 in 9-Dy. Colour code: as per Fig. 1

No frequency dependence is observed in ac magnetic susceptibility data for 9-Dy


measured in zero applied dc field. However, an external static dc field of 5000 Oe affords
a strong frequency dependence and peak maxima are evident in

m''

above 2K. The

Arrhenius analysis yields an energy barrier to the reversal of magnetisation Ueff/kb = 54.3
K (0 = 2 10-9 s).

Tetranuclear {Ln4} Compounds


Slow magnetic relaxation was observed for the Dy analogue of the series
(HDABCO)8H5Li8[Ln4As5W40O144(H2O)10(gly)2] (10) (DABCO = 1,4-diazabicyclooctane; gly =
zwitterionic glycine, Fig. 11).[47] Compounds (Ln = Gly, Dy, Tb and Ho) of series 10, are

24

obtained following the addition of DABCO to an acidic aqueous solution of the POM
precursor K14[As2W19O67(H2O)] and lanthanoid ions.

Fig. 11. Structural representations of the polyanion [Ln4As5W40O144(H2O)10(gly)2]21- in 10.


Colour code as per Fig. 1.

Unlike most of the other Ln-POM families discussed above, the four lanthanoid
ions in 10 are not sandwiched between two lacunary POM ligands. However, as is
observed for 1 and 2, the two distinct lanthanoid ions exhibit approximate trigonal
dodecahedral and square antiprismatic coordination geometries.[34] The shortest
intramolecular LnLn separation is 5.91 for 10-Dy.
Alternating current magnetic susceptibility measurements in the absence of an
applied dc field reveal out-of-phase components for 10-Dy only, although clear

m''.peak

maxima are not observed. Nevertheless, the energy barrier to the reversal of
magnetisation was determined by a generalised Deybe model. The Arrhenius fitting of
the temperature dependence of the relaxation time yielded a barrier of 3.9 K (0 = 1.9
10-5 s). In order to model the magnetic behaviour, the two crystallographically distinct
lanthanoid(III) ions were treated as a single centre with approximate D4d symmetry.

25

Under this assumption, the magnetic susceptibility data were fit to a simplified
Hamiltonian to yield energy levels for 10-Tb, 10-Dy and 10-Ho (Fig. 12).

Fig. 12. Energy levels and eigenfunctions composition of ground state multiplets for 10.
(This image is adapted from ref. 47 with permission. Copyright 2011 The American
Chemical Society).

Magnetostructural correlations

As is the case for Ln-SMMs with organic ligands,[6] intramolecular exchange interactions
between lanthanoid ions are of little importance to the magnetic behaviour exhibited by
poly-lanthanoid containing Ln-POM SMMs, the magnetic properties of which can thus be
treated as arising from single ion effects. Magnetostructural correlation remains a
challenge for lanthanoid-based SMMs.[6,48] Using the electrostatic argument of Rinehart
and Long it is possible to rationalise the behaviour of family 1 in which the slightly
distorted D4d pseudo-symmetry and the compressed geometry of the square antiprism

26

allows the POM ligands to be considered as "equatorial". The SMM behaviour of 1-Er can
then be ascribed to its prolate charge distribution, while the smaller energy barrier for 1Ho is due to its slightly oblate, almost spherical, quadrupole charge distribution. This
argument is consistent with the absence of slow relaxation for 1-Dy and 1-Tb, both with
oblate charge distributions in the presence of an equatorial geometry for the ligands.
This is in contrast to the first-reported Ln-based SMMs, the double-decker terbium and
dysprosium phthalocyanine complexes, for which the ligand arrangement is
geometrically neutral. The differences in the geometry of ligands between the POM and
phthalocyanine based Ln-SMMs are evident also from the crystal fields parameters
obtained from the fitting of susceptibility data (Fig. 13), which were performed using
approximate Hamiltonians in Stevens notation (Eqn. 1) for 1, 2, and [LnPc2]-, and in
Wybourne notation (Eqn. 2) for 10. [12,36,47]. The two formalisms are equivalent and the
parameters can be easily converted from one to another.

= 02 2 20 + 04 4 40 + 06 6 60

(1)

6
6
2
4
4
= 02

=1 0 () + 0 =1 0 () + 0 =1 0 ()

(2)

Negative values for 02 <r2> and 04 <r4> and positive values for 06 <r6> parameters
were determined for the Ln-POM complexes, indicating a compressed geometry in a
point charge model, whereas the value of 02 <r2> for [LnPc2]- is of opposite sign with
respect to that of the Ln-POM families, indicating a neutral disposition of charges. For
families 2 and 10, despite the high order D4d pseudo-symmetry, the greater distortion
from axial symmetry renders the oblate versus prolate model inapplicable. It in fact fails

27

completely for 2 where dysprosium(III) and holmium(III) (oblate ions) and erbium(III) and
ytterbium(III) (prolate ions) both exhibit SMM behaviour and also for 10, where the
dysprosium analogue is a SMM against expectations considering the compressed
geometry. The limits of this approach have been shown in a recent work where the
anisotropy axis of a Dy-SMM has been found both theoretically and experimentally to be
almost perpendicular to the highest pseudo 4-fold symmetry axis of the molecule.[49,50].

Fig. 13. Plot of crystal field parameters: 02 <r2> (squares), 04 <r4> (circles) and 06 <r6>
(triangles) determined for [LnPc2]- (green dashed line), 1 (black solid line), 2 (red solid
line) and 10 (blue dotted line).

The Rinehart and Long model is helpful to explain the change in magnetic
properties observed to accompany the reversible chemical switching between between
5-Dy2 and 6-Dy2.[43] The loss of acetone ligands from 7-coordinate dysprosium centres in

28

5-Dy2 to give 6-coordinate dysprosium in 6-Dy2 increases the local symmetry and reduces
inter-ligand repulsion. This stabilises the oblate shaped quadrupole charge distribution of
dysprosium(III), enhancing the anisotropy barrier to magnetisation reversal for 6-Dy2. For
the heterodinuclear family 6-DyLn, as the ionic radius of the second lanthanoid ions
increase, the distance between the two more negatively charged 2-hydroxo ligands also
increases, leading to an enhanced stabilisation of the oblate quadrupole charge
distribution of the dysprosium(III) ion and a larger energy barrier to magnetisation
reversal.[44] This series of compounds represents the first successful attempt to
systematically investigate the change in the energy barrier by varying the lanthanoid
partner ions in a heterodinuclear system. Fittingly, compound 6-DyEu has emerged from
this study with the highest energy barrier to magnetisation reversal determined for a LnPOM SMM.
Clearly more sophisticated theoretical models are required to fully elucidate the
correlation between structure and magnetic properties. Density functional theory and
CASSF calculations seem unavoidable, although the use of ab-initio methods in such
extended systems is a formidable task. Of relevance is a recent report of a simple
electrostatic model to describe the magnetic anisotropy in Dy-SMMs,[51] which is yet to
be applied to POM systems.

Outlook

Polyoxometalates are robust and versatile ligands for lanthanoid ions that have afforded
a number of Ln-SMMs. They are able to provide otherwise inaccessible coordination
geometries and crystal fields to lanthanoid ions, as well as an insulating diamagnetic

29

framework. Polyoxometalate-based Ln-SMMs share with other Ln-SMMs the common


issue of inadequate models for magnetostructural correlations. However, in addition to
the need for improved theoretical approaches, is the need for systematic studies of
isostructural analogues with different lanthanoid ions, with Ln-POMs well suited to
provide the requisite families of compounds.
Further advances in the field may involve surface functionalisation of POM ligands
in order to alter the nucleophilicity of the oxygen donor atoms that coordinate the
lanthanoid ions and so modulate the SMM properties, as has been reported for other LnSMMs.[52,53] The simultaneous coordination of both POM and organic ligands to
lanthanoid centres may also be pursued to access new coordination environments and
crystal fields. Varying the redox state of the coordinating ligands has proved important
for Ln-SMMs with organic ligands,[54,55] but is yet to be explored for Ln-POM SMMs, with
the rich redox chemistry of POMs very promising in this regard. Reduced, mixed-valence
POM ligands can potentially provide an additional source of spin to Ln-POM SMMs, or a
route to altering the crystal field and anisotropy. Notably, most of the Ln-POM SMMs
reported to date are comprised of polyoxotungstate ligands, with a very limited presence
of polyoxomolybdates and no polyoxovanadates. In order to make better use of the
structural versatility of polyoxometalates, the study of Ln-POM SMMs must be extended
beyond polyoxotungstates.

References

[1]

D. Gatteschi, R. Sessoli, Angew. Chem. Int. Ed. 2003, 42, 268. doi:
10.1002/anie.200390099

30

[2]

R. E. P. Winpenny, Molecular Cluster Magnets (World Scientific Series in


Nanoscience and Nanotechnology), World Scientific Publishing Company, 2011.

[3]

D. Gatteschi, R. Sessoli, J. Villain, Molecular Nanomagnets (Mesoscopic Physics and


Nanotechnology), Oxford University Press, USA, 2011.

[4]

D. Stepanenko, M. Trif, D. Loss, Inorg. Chim. Acta 2008, 361, 3740. doi:
10.1016/j.ica.2008.02.066

[5]

N. Ishikawa, M. Sugita, W. Wernsdorfer, Angew. Chem. Int. Ed. 2005, 44, 2931. doi:
10.1002/anie.200462638

[6]

D. N. Woodruff, R. E. P. Winpenny, R. A. Layfield, Chem. Rev. 2013, 113 (7), 5110.


doi: 10.1021/cr400018q

[7]

C. J. Milios, A. Vinslava, W. Wernsdorfer, S. Moggach, S. Parsons, S. P. Perlepes, G.


Christou, E. K. Brechin, J. Am. Chem. Soc. 2007, 129, 2754. doi: 10.1021/ja068961m

[8]

D. Yoshihara, S. Karasawa, N. Koga, J. Am. Chem. Soc. 2008, 130, 10460. doi:
10.1021/ja802895d

[9]

O. Waldmann, Inorg. Chem. 2007, 46, 10035. doi: 10.1021/ic701365t

[10]

J. M. Zadrozny, D. J. Xiao, M. Atanasov, G. J. Long, F. Grandjean, F. Neese, J. R.


Long, Nat. Chem. 2013, 5, 577. doi: 10.1038/nchem.1630

[11]

N. Ishikawa, M. Sugita, T. Ishikawa, S.-Y. Koshihara, Y. Kaizu, J. Am. Chem. Soc.


2003, 125, 8694. doi: 10.1021/ja029629n

[12]

N. Ishikawa, M. Sugita, T. Okubo, N. Tanaka, T. Iino, Y. Kaizu, Inorg. Chem. 2003,


42, 2440. doi: 10.1021/ic026295u

[13]

C. R. Ganivet, B. Ballesteros, G. de la Torre, J. M. Clemente-Juan, E. Coronado, T.


Torres, Chem. - A Eur. J. 2013, 19, 1457. doi: 10.1002/chem.201202600

[14]

R. Sessoli, A. K. Powell, Coord. Chem. Rev. 2009, 253, 2328. doi:


10.1016/j.ccr.2008.12.014

[15]

L. Sorace, C. Benelli, D. Gatteschi, Chem. Soc. Rev. 2011, 40, 3092. doi:
10.1039/c0cs00185f

[16]

R. A. Layfield, Organometallics 2014, 33, 1084. doi: 10.1021/om401107f

[17]

J. M. Clemente-Juan, E. Coronado, A. Gaita-Ario, Chem. Soc. Rev. 2012, 41, 7464.


doi: 10.1039/c2cs35205b

[18]

J. Luzon, R. Sessoli, Dalton Trans. 2012, 41, 13556. doi: 10.1039/c2dt31388j

31

[19]

J. D. Rinehart, J. R. Long, Chem. Sci. 2011, 2, 2078. doi: 10.1039/c1sc00513h

[20]

P. Zhang, L. Zhang, C. Wang, S. Xue, S.-Y. Lin, J. Tang, J. Am. Chem. Soc. 2014, DOI
10.1021/ja500793x

[21]

K. Meihaus, J. Long, J. Am. Chem. Soc. 2013, 135, 1795217957. doi:


10.1021/ja4094814

[22]

M. Maeda, S. Hino, K. Yamashita, Y. Kataoka, M. Nakano, T. Yamamura, T.


Kajiwara, Dalton Trans. 2012, 41, 13640. doi: 10.1039/c2dt31399e

[23]

N. F. Chilton, S. K. Langley, B. Moubaraki, A. Soncini, S. R. Batten, K. S. Murray,


Chem. Sci. 2013, 4, 1719. doi: 10.1039/c3sc22300k

[24]

M. T. Pope, Heteropoly and Isopoly Oxometalates, Springer-Verlag, 1983.

[25]

D.-L. Long & L. Cronin (Eds.), Dalton Trans. 2012, 41 (issue 33), 9799-10106 entire
issue. doi: 10.1039/c2dt90121h

[26]

T. Liu & U. Kortz (Eds.), Eur. J. Inorg. Chem. 2013, (issues 10-11), 1556-1967 entire
issue. doi: 10.1002/ejic.201300230

[27]

X. Lpez, J. J. Carb, C. Bo, J. M. Poblet, Chem. Soc. Rev. 2012, 41, 7537. doi:
10.1039/c2cs35168d

[28]

T. Yamase, Chem. Rev. 1998, 98, 307. doi: 10.1021/cr9604043

[29]

C. M. Granadeiro, B. de Castro, S. S. Balula, L. Cunha-Silva, Polyhedron 2013, 52,


10. doi: 10.1016/j.poly.2012.09.037

[30]

C. Ritchie, C. Boskovic, in Polyoxometalate Chem. Some Recent Trends (Ed.: F.


Scheresse), World Scientific, Singapore, 2013, 177.

[31]

C. Boglio, G. Lemire, B. Hasenknopf, S. Thorimbert, E. Lacte, M. Malacria,


Angew. Chem. Int. Ed. 2006, 45, 3324. doi: 10.1002/anie.200600364

[32]

C. Ritchie, E. G. Moore, M. Speldrich, P. Kgerler, C. Boskovic, Angew. Chem. Int.


Ed. 2010, 49, 7702. doi: 10.1002/anie.201002320

[33]

C. Ritchie, V. Baslon, E. G. Moore, C. Reber, C. Boskovic, Inorg. Chem. 2012, 51,


1142. doi: 10.1021/ic202349u.

[34]

M. Llunell, D. Casanova, J. Cirera, J. M. Bofill, P. Alemany, S. Alvarez, M. Pinsky, D.


Avnir, SHAPE 2.0, Universitat De Barcelona And The Hebrew University Of
Jerusalem, Barcelona, 2003.

[35]

M. A. Aldamen, J. M. Clemente-Juan, E. Coronado, C. Mart-Gastaldo, A. GaitaArio, J. Am. Chem. Soc. 2008, 130, 8874. doi: 10.1021/ja801659m

32

[36]

M. A. AlDamen, S. Cardona-Serra, J. M. Clemente-Juan, E. Coronado, A. GaitaArio, C. Mart-Gastaldo, F. Luis, O. Montero, Inorg. Chem. 2009, 48, 3467. doi:
10.1021/ic801630z

[37]

S. Ghosh, S. Datta, L. Friend, S. Cardona-Serra, A. Gaita-Ario, E. Coronado, S. Hill,


Dalton Trans. 2012, 41, 13697. doi: 10.1039/c2dt31674a

[38]

S. Cardona-Serra, J. M. Clemente-Juan, E. Coronado, A. Gaita-Ario, A. Camn, M.


Evangelisti, F. Luis, M. J. Martnez-Prez, J. Ses, J. Am. Chem. Soc. 2012, 134,
14982. doi: 10.1021/ja305163t

[39]

R. D. Peacock, T. J. R. Weakley, in J. Chem. Soc. A, The Royal Society Of Chemistry,


1971, 1836. doi: 10.1039/j19710001836

[40]

Y. Jeannin, J. Martin-Frre, D. J. Choi, M. T. Pope, in Inorg. Synth. (Ed.: A.P.


Ginsberg), John Wiley & Sons, Inc, Hoboken, NJ, USA., 1990, 115. doi:
10.1002/9780470132586.ch20

[41]

I. Creaser, M. C. Heckel, R. J. Neitz, M. T. Pope, Inorg. Chem. 1993, 32, 1573. doi:
10.1021/ic00061a010

[42]

L. Liu, F. Li, L. Xu, X. Liu, G. Gao, J. Solid State Chem. 2010, 183, 350. doi:
10.1016/j.jssc.2009.11.013

[43]

K. Suzuki, R. Sato, N. Mizuno, Chem. Sci. 2013, 4, 596. doi: 10.1039/c2sc21619a

[44]

R. Sato, K. Suzuki, M. Sugawa, N. Mizuno, Chem. - A Eur. J. 2013, 19, 12982. doi:
10.1002/chem.201302596

[45]

X. Feng, W. Zhou, Y. Li, H. Ke, J. Tang, R. Clrac, Y. Wang, Z. Su, E. Wang, Inorg.
Chem. 2012, 51, 2722. doi: 10.1021/ic202418y

[46]

F. Li, W. Guo, L. Xu, L. Ma, Y. Wang, Dalton Trans. 2012, 41, 9220. doi:
10.1039/c2dt12277d

[47]

C. Ritchie, M. Speldrich, R. W. Gable, L. Sorace, P. Kgerler, C. Boskovic, Inorg.


Chem. 2011, 50, 7004. doi: 10.1021/ic200366a

[48]

P. Zhang, Y.-N. Guo, J. Tang, Coord. Chem. Rev. 2013, 257, 1728. doi:
10.1016/j.ccr.2013.01.012

[49]

M.-E. Boulon, G. Cucinotta, J. Luzon, C. DeglInnocenti, M. Perfetti, K. Bernot, G.


Calvez, A. Caneschi, R. Sessoli, Angew. Chem. Int. Ed. 2013, 52, 350. doi:
10.1002/anie.201205938

[50]

G. Cucinotta, M. Perfetti, J. Luzon, M. Etienne, P.-E. Car, A. Caneschi, G. Calvez, K.


Bernot, R. Sessoli, Angew. Chem. Int. Ed. 2012, 124, 1638. doi:
10.1002/ange.201107453

33

[51]

N. F. Chilton, D. Collison, E. J. L. McInnes, R. E. P. Winpenny, A. Soncini, Nat.


Commun. 2013, 4, 2551. doi:10.1038/ncomms3551

[52]

F. Habib, G. Brunet, V. Vieru, I. Korobkov, L. F. Chibotaru, M. Murugesu, J. Am.


Chem. Soc. 2013, 135, 13242. doi: 10.1021/ja404846s

[53]

M. Gonidec, D. B. Amabilino, J. Veciana, Dalton Trans. 2012, 41, 13632. doi:


10.1039/c2dt31171b

[54]

N. Ishikawa, M. Sugita, N. Tanaka, T. Ishikawa, S. Koshihara, Y. Kaizu, Inorg. Chem.


2004, 43, 5498. doi: 10.1021/ic049348b

[55]

S. Takamatsu, T. Ishikawa, S. Koshihara, N. Ishikawa, Inorg. Chem. 2007, 46, 7250.


doi: 10.1021/ic700954t

34

Entry for Table of Contents

Lanthanoid-based single-molecule magnets (Ln-SMMs) hold clear promise for the


achievement of large energy barriers to magnetisation reversal. The structural and
electronic versatility of polyoxometalates renders them a useful class of supporting
ligands for Ln-SMMs. This review discusses Ln-SMMs with polyoxometalate ligands,
focusing on structure, magnetic properties and magnetostructural correlations.

You might also like